Tin-Free Radical Addition of Acyloxymethyl to Imines - ACS Publications

Jul 31, 2008 - Ken-ichi Yamada , Masaru Maekawa , Tito Akindele , Mayu Nakano , Yasutomo Yamamoto and Kiyoshi Tomioka. The Journal of Organic ...
1 downloads 0 Views 97KB Size
ORGANIC LETTERS

Tin-Free Radical Addition of Acyloxymethyl to Imines

2008 Vol. 10, No. 17 3805-3808

Ken-ichi Yamada, Mayu Nakano, Masaru Maekawa, Tito Akindele, and Kiyoshi Tomioka* Graduate School of Pharmaceutical Sciences, Kyoto UniVersity, Yoshida, Sakyo-ku, Kyoto 606-8501, Japan [email protected] Received July 2, 2008

ABSTRACT

Acyloxymethyl radicals were generated from the corresponding iodomethyl esters and successfully underwent addition to the CdN bond of N-Ts, N-PMP, and N-Dpp imines by the action of dimethylzinc or triethylborane. Ethyl acetate, toluene, and benzene as well as dichloromethane were suitable solvents. The utility of acyloxymethyl radicals as a hydroxymethyl anion equivalent was highlighted by the facile hydrolysis of the acyloxy moiety of the adducts to give the corresponding amino alcohol derivatives in good to high yield.

Synthetically useful transformations that are otherwise difficult to accomplish have been achieved through reactions involving radical species.1 In addition, these reactions are usually carried out under mild conditions. An R-oxygenated alkyl radical is classified as a stabilized nucleophilic radical and is accordingly an R-oxygenated carbanion equivalent, while the anion counterpart, R-alkoxycarbanion, is a relatively unstable species. Examples of the most successful methods to generate R-oxygenated carbanion are tin-lithium exchange of (1-alkoxyalkyl)tin compounds2 and iodinemagnesium exchange of iodomethyl carboxylates.3 Usually, strictly anhydrous and deoxygenated atmosphere, strongly basic conditions, and also low temperature are required for this type of carbanion chemistry. In contrast, R-alkoxy and R-hydroxyalkyl radicals have been generated via C-H bond cleavage under milder conditions and utilized in C-C bond forming reactions as nucleophilic R-oxygenated carbon units.4 The major problem of this strategy was, however, the requirement of a large amount (usually 20-250 equiv) of the corresponding ether or alcohol as a radical source. (1) (a) Radicals in Organic Synthesis; Renaud, P., Sibi, M. P., Eds.; Wiley-VCH: Weinheim, 2001; Vols. 1 and 2. (b) Radicals in Synthesis I & II; Gansa¨uer, A., Ed.; Topics in Current Chemistry 263 and 264; SpringerVerlag: Berlin, Heidelberg, 2006. (2) Still, W. C.; Mitra, A. J. Am. Chem. Soc. 1978, 100, 1927–1928. (3) Avolio, S.; Malan, C.; Marek, I.; Knochel, P. Synlett 1999, 1820– 1822. 10.1021/ol8014978 CCC: $40.75 Published on Web 07/31/2008

 2008 American Chemical Society

The generation of this class of radical via C-heteroatom bond cleavage using a tin compound has been utilized in sugar chemistry as well as in a number of total syntheses.5 Although R-oxygenated C-centered radicals were efficiently generated from the radical sources in these examples, the use of organotinhydrides is a critical drawback, especially from an ecological point of view, because of difficulty in complete removal and toxicity of tin compounds.6 Moreover, the radical source, R-halogenated ether is rather unstable and easily hydrolyzed under ordinary atmosphere. R-Acyloxyalkyl radical is also expected to react as a nucleophilic R-oxygenated carbon unit. The corresponding radical source, R-halogenated alkyl benzoate and pivalate are so stable (4) (a) Fraser-Reid, B.; Anderson, R. C.; Hicks, D. R.; Walker, D. L. Can. J. Chem. 1977, 55, 3986–3995. (b) Fontana, F.; Minisci, F.; Yan, Y. M.; Zhao, L. Tetrahedron Lett. 1993, 34, 2517–2520. (c) Gong, J.; Fuchs, P. L. J. Am. Chem. Soc. 1996, 118, 4486–4487. (d) Peleta, O.; Cı´rkva, V.; Buskova´, Z.; Bo¨hm, S. J. Fluorine Chem. 1997, 86, 155–171. (e) Kim, S.; Kim, N.; Chung, W.-J.; Cho, C. H. Synlett 2001, 937–940. (f) Torrente, S.; Alonso, R. Org. Lett. 2001, 3, 1985–1987. (g) Jang, Y.-J.; Shih, Y.-K.; Liu, J.-Y.; Kuo, W.-Y.; Yao, C.-F. Chem. Eur. J. 2003, 9, 2123–2128. (h) Ferna´ndez, M.; Alonso, R. Org. Lett. 2003, 5, 2461–2464. (i) Ferna´ndez, M.; Alonso, R. Org. Lett. 2005, 7, 11–14. (j) Oka, R.; Nakayama, M.; Sakaguchi, S.; Ishii, Y. Chem. Lett. 2006, 35, 1104–1105. (k) Geraghty, N. W. A.; Lally, A. Chem. Commun. 2006, 4300–4302. (5) (a) Lee, E. In Radicals in Organic Synthesis; Renaud, P., Sibi, M. P., Eds.; Wiley-VCH: Weinheim, 2001; Vol. 2, pp 303-333. (b) Buckmelter, A. J.; Rychnovsky, S. D. In Radicals in Organic Synthesis; Renaud, P., Sibi, M. P., Eds.; Wiley-VCH: Weinheim, 2001; Vol. 2, pp 334-349. (6) Review: Neumann, W. P. Synthesis 1987, 665–683.

toward hydrolysis7 that the purification by silica gel column chromatography is possible. However, the synthetic utility of R-acyloxyalkyl radical has scarcely been studied. Besides, reported examples are limited to addition to an intramolecular CdC bond,8,9 with the exception of one example of intermolecular addition of propanoyloxymethyl radical, generated from iodomethyl propanoate using an organotin reagent.10,11 We have studied the dimethylzinc-air-initiated direct generation of R-alkoxyalkyl radicals from ethers through hydrogen abstraction and their reactions for the functionalization of ethers.12 We also succeeded in the tin-free generation of primary alkyl radicals from the corresponding alkyl iodides via dimethylzinc-air-radical process.13-15 Herein, we report the tin-free generation of acyloxymethyl radicals from the corresponding iodomethyl esters and their reaction with imines to show their utility as hydroxymethyl anion equivalents. First, we examined the reaction of iodomethyl pivalate (1a)16 with N-tosylbenzaldimine (2a). To a solution of 1a (3 mmol) and 2a (1 mmol) in dichloromethane (1 mL) was added a 1.0 M solution of dimethylzinc in hexane (1 mL, 1 mmol). The mixture was stirred at room temperature under ordinary atmosphere, while a solution of dimethylzinc (1 mL, 1 mmol) was added every 2 h. After addition of a total amount of 5 mmol of dimethylzinc, the mixture was stirred (7) Ulich, L. H.; Adams, R. J. Am. Chem. Soc. 1921, 43, 660–667. (8) Beckwith, A. L. J.; Pigou, P. E. J. Chem. Soc., Chem. Commun. 1986, 85–86. (9) Meyer, C.; Marek, I.; Courtemanche, G.; Normant, J.-F. Tetrahedron 1994, 50, 11665–11692. (10) Degueil-Castaing, M.; Navarro, C.; Ramon, F.; Maillard, B. Aust. J. Chem. 1995, 48, 233–240. (11) Recently, intermolecular addition of phthalimidylmethyl radical to electron-rich CdC bond has been reported: Quiclet-Sire, B.; Zard, S. Z. Org. Lett. 2008, 10, 3279–3282. (12) (a) Yamada, K.; Fujihara, H.; Yamamoto, Y.; Miwa, Y.; Taga, T.; Tomioka, K. Org. Lett. 2002, 4, 3509–3511. (b) Yamada, K.; Yamamoto, Y.; Tomioka, K. Org. Lett. 2003, 5, 1797–1799. (c) Yamada, K.; Yamamoto, Y.; Maekawa, M.; Tomioka, K. J. Org. Chem. 2004, 64, 1531–1534. (d) Yamamoto, Y.; Yamada, K.; Tomioka, K. Tetrahedron Lett. 2004, 45, 795– 797. (e) Yamamoto, Y.; Maekawa, M.; Akindele, T.; Yamada, K.; Tomioka, K. Tetrahedron 2005, 61, 379–384. (f) Akindele, T.; Yamamoto, Y.; Maekawa, M.; Umeki, H.; Yamada, K.; Tomioka, K. Org. Lett. 2006, 8, 5729–5732. (g) Yamada, K.; Umeki, H.; Maekawa, M.; Yamamoto, Y.; Akindele, T.; Nakano, M.; Tomioka, K. Tetrahedron 2008, 64, 7258–7265. (h) Review: (i) Yamada, K.; Yamamoto, Y.; Tomioka, K. J. Synth. Org. Chem. Jpn. 2004, 62, 1158–1165. (13) Yamada, K.; Yamamoto, Y.; Maekawa, M.; Akindele, T.; Umeki, H.; Tomioka, K. Org. Lett. 2006, 8, 87–89. (14) Studies on reactions of dialkylzinc and molecular oxygen: (a) Lewinski, J.; Marciniak, W.; Lipkowski, J.; Justyniak, I. J. Am. Chem. Soc. 2003, 125, 12698–12699. (b) Lewinski, J.; Sliwinski, W.; Dranka, M.; Justyniak, I.; Lipkowski, J. Angew. Chem., Int. Ed. 2006, 45, 4826–4829. (c) Jana, S.; Berger, R. J. F.; Fro¨hlich, R.; Pape, T.; Mitzel, N. W. Inorg. Chem. 2007, 46, 4293–4297. (15) For dialkylzinc-mediated radical reactions, see: Bazin, S.; Feray, L.; Bertrand, M. P. Chimia 2006, 60, 260–265, and references cited therein. (16) Prepared from the corresponding commercially available chloride according to the reported procedure: Knochel, P.; Chou, T.; Jubert, C.; Rajagopal, D. J. Org. Chem. 1993, 58, 588–599. (17) Reviews: (a) Olliver, C.; Renaud, P. Chem. ReV. 2001, 101, 3415– 3434. (b) Yorimitsu, H.; Oshima, K. In Radicals in Organic Synthesis; Renaud, P., Sibi, M. P., Eds.; Wiley-VCH: Weinheim, 2001; Vol. 1, pp 11-27. (18) The other reactions of methoxymethyl radical, generated from methyl iodomethyl ether, were also conducted at -78 °C: (a) Halland, N.; Jørgensen, A. J. Chem. Soc., Perkin Trans. 1 2001, 1290–1295. (b) Bazin, S.; Feray, L.; Vanthuyne, N.; Bertrand, M. P. Tetrahedron 2007, 63, 77– 85. 3806

for another 6 h at room temperature to give adduct 3a in 66% yield (Scheme 1).

Scheme 1

The reaction using diethylzinc, in place of dimethylzinc, failed to give the expected product 3a but instead the ethyl adduct in 78% yield along with the reduced product, N-benzyltoluenesulfonamide, in 9% yield. When a hexane solution of triethylborane17 was used instead of dimethylzinc, the reaction proceeded more cleanly to give 3a in 96% yield after 20 h (Table 1, entry 1). The ester functionality of 1

Table 1. Addition of Acyloxymethyl Radicals, Generated from 1, to Imines 2 Mediated by Et3B-Aira

entry

1/R1

2

R2

R3

1 2 3 4 5b 6 7 8d 9

1a/t-Bu 1b/Ph 1a/t-Bu 1b/Ph 1a/t-Bu 1a/t-Bu 1a/t-Buc 1a/t-Bu 1a/t-Bu

2a 2a 2b 2b 2c 2d 2e 2f 2g

Ph Ph Ph Ph Ph 4-ClC6H4 4-MeC6H4 4-MeOC6H4 Ph(CH2)2

4-tolSO2 4-tolSO2 4-MeOC6H4 4-MeOC6H4 Ph2PO 4-tolSO2 4-tolSO2 4-tolSO2 4-tolSO2

Et3B time 3/yield (equiv) (h) (%) 9 5 5 5 20 8 18 9 8

20 14 7 8 48 20 48 22 24

3a/96 3b/79 3c/89 3d/84 3e/80 3f/88 3g/82 3h/94 3i/67

a

1 (3 equiv) and Et3B (3 equiv) were initially added, and the rest of Et3B was portionwise added (1 equiv/2 h). b Toluene was used as a solvent in place of CH2Cl2. BF3·OEt2 (total 6 equiv) was added in four portions. c 9 equiv. d BF3·OEt2 (total 3 equiv) was added in three portions.

and 3 was inert under these conditions. It is worthy of note that the reaction can be conducted at room temperature under ordinary atmosphere. In contrast, the reaction of methoxymethyl radical, generated from methyl iodomethyl ether with dimethylzinc-air, required -78 °C to give the methoxymethylated analogue of 3a in 95% yield, and the reaction at room temperature resulted in a complex mixture including no desired adduct.18 Iodomethyl benzoate (1b)19 was also a good radical precursor to give benzoyloxymethylated adduct 3b20 in 79% (19) Prepared from the corresponding commercially available chloride according to the reported procedure: Wittig, G.; Jautelat, M. Liebigs Ann. Chem. 1967, 702, 24–37. (20) Sun, X.; Ye, S.; Wu, J. Eur. J. Org. Chem. 2006, 21, 4787–4790. Org. Lett., Vol. 10, No. 17, 2008

yield after 14 h under similar conditions (entry 2). The reaction proceeded smoothly with N-(4-methoxyphenyl) (PMP) imine 2b, and products 3c and 3d were obtained in 89% and 84% yields after 7 and 8 h, respectively (entries 3 and 4). Interestingly, the reactions of N-PMP imines were faster than those with more electron-deficient N-tosyl imines. The reaction of N-diphenylphosphinoyl (Dpp) imine 2c also proceeded in toluene by using 20 equiv of triethylborane to give 3e in 80% yield in the presence of trifluoroborane diethyl etherate (entry 5). In the absence of trifluoroborane diethyl etherate, the reaction produced 3e in only 20% yield and 2c was recovered in 40% yield after 48 h using 20 equiv of triethylborane. The products were also obtained in good yields in the reaction of N-tosyl imines 2d and 2e derived from 4-chloro and 4-methylbenzaldehyde (entries 6 and 7), although the latter required 9 equiv of 1a and prolonged reaction time. The presence of trifluoroborane also facilitated the reaction of electron-rich 4-methoxybenzaldehyde imine 2f. In the presence of trifluoroborane diethyl etherate (3 equiv), adduct 3h was obtained in 94% yield after 22 h (entry 8). In the absence of trifluoroborane diethyl etherate, 3h was obtained in only 41% yield along with recovered 2f in 45% yield after 60 h using 9 equiv of 1a and 20 equiv of triethylborane (Scheme 2). Catalytic amounts (0.2 equiv) of other additives,

Scheme 3

resulting in quantitative recovery of the starting aldehyde. This result encouraged us to perform a three-component reaction. To a mixture of 1a (3 equiv), benzaldehyde, and p-anisidine (1 equiv) was added triethylborane (10 equiv in eight portions). Addition of pivaloyloxymethyl radical exclusively took place to imine 2b, generated in situ, to give adduct 3c in 65% yield after 22 h (Scheme 4). No addition product to the CdO bond of benzaldehyde was detected.

Scheme 2 Scheme 4

Demasking of the alcohol moiety of pivalate and benzoate product 3 was readily achieved by basic hydrolysis with potassium hydroxide in aqueous methanol at room temperature to give the corresponding amino alcohols 421 in good to high yield (Scheme 5). copper(II) triflate and scandium triflate also promoted the reaction to give 3h in 72% and 65% yields after 22 and 7 h, respectively, using 9 equiv of triethylborane. On the contrary, bismuth triflate and indium triflate failed to facilitate the reaction to give 3h in 36% and